Next Article in Journal
Targeting the CD40-CD154 Signaling Pathway for Treatment of Autoimmune Arthritis
Next Article in Special Issue
Identification of Cancer Stem Cell Subpopulations in Head and Neck Metastatic Malignant Melanoma
Previous Article in Journal
The uPAR System as a Potential Therapeutic Target in the Diseased Eye
Previous Article in Special Issue
Role of miRNA-Regulated Cancer Stem Cells in the Pathogenesis of Human Malignancies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cancer Stem Cells and Targeting Strategies

1
The John van Geest Cancer Research Centre, School of Science and Technology, Nottingham Trent University, Clifton Lane, Nottingham NG11 8NS, UK
2
Department of Precision Medicine, University of Campania “Luigi Vanvitelli”, 80138 Naples, Italy
*
Author to whom correspondence should be addressed.
Cells 2019, 8(8), 926; https://doi.org/10.3390/cells8080926
Submission received: 5 July 2019 / Revised: 5 August 2019 / Accepted: 15 August 2019 / Published: 18 August 2019
(This article belongs to the Special Issue CSCs Identification and Targeting Therapies: Advances and Challenges)

Abstract

:
Chemoresistance is a major problem in cancer therapy as cancer cells develop mechanisms that counteract the effect of chemotherapeutic compounds, leading to relapse and the development of more aggressive cancers that contribute to poor prognosis and survival rates of treated patients. Cancer stem cells (CSCs) play a key role in this event. Apart from their slow proliferative property, CSCs have developed a range of cellular processes that involve drug efflux, drug enzymatic inactivation and other mechanisms. In addition, the microenvironment where CSCs evolve (CSC niche), effectively contributes to their role in cancer initiation, progression and chemoresistance. In the CSC niche, immune cells, mesenchymal stem cells (MSCs), endothelial cells and cancer associated fibroblasts (CAFs) contribute to the maintenance of CSC malignancy via the secretion of factors that promote cancer progression and resistance to chemotherapy. Due to these factors that hinder successful cancer therapies, CSCs are a subject of intense research that aims at better understanding of CSC behaviour and at developing efficient targeting therapies. In this review, we provide an overview of cancer stem cells, their role in cancer initiation, progression and chemoresistance, and discuss the progress that has been made in the development of CSC targeted therapies.

Graphical Abstract

1. Introduction

In 2018, the World Health Organization (WHO) asserted that cancer is responsible for the death of 9.6 million people worldwide [1]. The data also suggest that despite the availability of different drugs, their efficacy requires further improvement with a focus on cancer complexity. In this intricate system, cancer stem cells play a crucial role, as they mediate chemoresistance and cancer recurrence. It is now clear that cancer stem cells (CSCs) represent a small population of cancer cells within a tumour and that they have typical characteristics related to stem cells. Like their stem cell counterparts, CSCs possess ‘stemness’ properties which are reflected in their capacity to self-renew and to generate differentiated cells that contribute to tumour heterogeneity. CSCs appear to be generated from mutations affecting adult stem cells that are at the source of organogenesis and tissue homeostasis [2,3,4]. Although debates about their origin persist, there is an urgent and continuous need to better understand the function of CSCs in cancer initiation and progression and, more importantly, develop CSC-specific targeting strategies.
Chemotherapy induces apoptosis by damaging DNA and/or inhibiting mitotic division. However, this therapeutic option is valid for highly dividing cancer cells and therefore has the potential to “miss out” slow dividing cells, leading to cancer relapse and recurrence in treated patients. Unfortunately, follow up chemotherapy that attempted to destroy refractory cancers cells has been mostly unsuccessful and was associated with a decrease in survival rates of treated cancer patients [5,6]. Due to their slow rate of division, CSCs mediate chemoresistance and recurrence following chemotherapy; however, additional mechanisms that confer CSCs a survival advantage include the expression of the ATP-binding cassette (ABC) transporters, a higher expression of aldehyde dehydrogenases (ALDHs), an increased resistance to apoptosis, and a timely activation of the DNA damage sensor and repair machinery [7]. Other mechanisms, that are not discussed in this review, involve epigenetic factors that play a role in regulating gene expression at the transcriptional or post-transcriptional levels [8,9]. All these properties and their clinical implications point to the importance of investigating and targeting CSCs.
It is well known that the microenvironment or niche is essential for stemness maintenance, and this also applies to CSCs where cell‒cell interactions within the niche are required for CSCs’ role in cancer initiation and progression [10]. The CSC niche is a complex environment that contains different types of cells, such as cancer-associated fibroblasts (CAFs), endothelial and immune cells, and mesenchymal stem cells (MSCs). These cells communicate through secreted growth factors and cytokines to promote tumorigenesis, angiogenesis, invasion and metastasis. The microenvironment of CSCs also contributes to CSCs’ resistance to drugs and other cancer therapies, which promotes cancer recurrence [11]. Several therapeutic strategies have been developed to target CSCs with the aim of preventing cancer recurrence. Although progress has been made through the development of targeting strategies, their therapeutic efficacy remains insufficient due to the absence of specific targetable biomarkers. In this review, we provide an overview of CSCs, their role in cancer initiation and progression, and discuss the targeting strategies that are being developed to target these highly aggressive cancer cells.

2. Cancer Stem Cells

“Omnis cellula e cellula” (“every cell stem from another cell”), with those words Rudolf Virchow, a German pathologist, gave the base of the cellular pathology. He had already postulated that the disease is due to changes happening in normal cells and that a tumour originates from immature cells [12,13]. However, the term cancer stem cell was coined by Sajiro Makino through the identification of a subpopulation of cells that were resistant to chemotherapy and with genetic characteristics that were different from the Bulk [13]. It was through studies carried out using the in vitro colony formation assay that Makino theory could be confirmed, and that suggested that a tumour could have originated from cells that had characteristics attributable to stem cells [14]. Further confirmation of Virchow’s idea came later from experiments using acute myeloid leukaemia (AML) cells that led to the identification of a population of cells with a stem-like phenotype [15]. The authors identified a population of CD34+CD38 AML stem cells that were capable of initiating AML in severe combined immune-deficient (SCID) mice. Although using a different set of markers, populations of CSCs were also found in other types of cancerous tissues, which confirmed their existence within other types of tumours and the role that they play in cancer initiation and progression [16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. Two main hypotheses have been proposed to explain how CSCs are originated within a tissue: the stochastic and the hierarchical models [36,37]. The stochastic model supports the hypothesis that any cell subjected to a specific set of somatic mutations could lead to the acquisition of self-renewal and differentiation properties that could contribute to tumour heterogeneity. On the other hand, the hierarchical model proposes that somatic mutations target specifically stem and progenitor cells that are hierarchically organized, resulting in the generation of self-renewing cancer stem and progenitor cells that produce differentiated progenies and recreate heterogenous tumours [38]. In summary, CSCs are populations of cancer cells that have the stem cell properties of self-renewal and ability to generate differentiated progenies.

3. Cancer Stemness, Therapy Resistance and Cancer Recurrence

A major challenge in cancer therapy is associated with chemoresistance and recurrence following chemotherapy [2,39]. In most cases, this resistance is associated with the presence, within tumours, of aggressive cancer cell populations that developed mechanisms of chemoresistance responsible for unsuccessful therapies and leading to decreased survival rates of treated cancer patients (Figure 1A) [7,10]. Several cancer studies identified these aggressive cancer populations as cancer stem cells (CSCs) (Table 1) [15,16,40,41,42,43,44,45,46]. CSC-mediated chemoresistance relies on several cellular mechanisms that include lower proliferation rates and the ability to intracellularly and enzymatically inactivate chemotherapeutic compounds (Figure 1B) [7].

3.1. Low Proliferation Rate

Chemotherapy targets highly proliferative cells through DNA damage and inhibition of mitotic division; however, its action is limited when it applies to slow and non-dividing cancer cells such as CSCs [48,49,50,51]. Indeed, several studies have shown that CSCs, from different tissue origin, could resist the effect of a wide range of drugs such as doxorubicin, temozolomide, cisplatin, paclitaxel, etoposide and methotrexate [52]. Hence, following chemotherapy, which destroys the bulk of the tumour, CSCs remain on-site and can re-initiate tumorigenic and metastatic processes. Hence, developing therapies that target dormant or slow dividing cells is essential to prevent recurrent cancers.

3.2. Expression of ATP-Binding Cassette (ABC) Transporters

This large superfamily of membrane proteins functions in the transmembrane transport of various substrates by converting energy obtained from ATP hydrolysis [53]. Although the type of ABC transporters varies from one type of cancer to another, CSCs express high levels of these transporters, which contributes to an increased transport of chemotherapeutic compounds out of the cells, contributing therefore to CSC chemoresistance. Some ABC transporters are involved in the transport of a large number of drugs, while others transport only a few. For instance, the ABCC1 (ATP-binding cassette transporter C1) transporter also known as MRP1 (Multi Drug Resistance Protein 1), expressed by CSCs in glioblastoma, is involved in the efflux of a variety of therapeutic compounds including methotrexate, edatrexate, ZD1694, doxorubicin, daunorubicin, epirubicin, idarubicin, etoposide, vincristine, vinblastine, paclitaxel, irinotecan, SN-38, flutamide and hydroxyflutamide [54,55,56]. ABCB1 (ATP-binding cassette transporter B1), also known as MDR1 (Multidrug Resistance Protein 1), expressed by CSCs found in ovaries and breast cancers, acute myeloid leukaemia (AML), glioblastoma and renal cell carcinoma, provides drug multi-resistance to anthracyclines actinomycin D, colchicine, etoposide, teniposide, methotrexate, mitomycin C, mitoxantrone, paclitaxel, docetaxel, vincristine and vinblastine [57,58,59,60]. Other ABC transporters have a limited drug transport capacity, such as ABCA1 (ATP-binding cassette transporter A1), which is expressed by serous ovarian CSCs and which transports cisplatin [59]. In summary, ABCs are drug specific transporters that may require combination therapies that use inhibitors for the different transporters.

3.3. Increased Expression of Aldehyde Dehydrogenases (ALDHs)

This superfamily of enzymes is crucial for the detoxification of endogenous and exogenous aldehyde substrates by catalysing the oxidation of aldehydes to carboxylic acids [61,62,63]. They are highly expressed in CSCs from different types of cancers and their elevated levels of expression correlate with worse prognosis in cancer patients [64,65,66]. Although the mechanisms by which they promote CSC chemoresistance is not well known, this activity may involve their role in retinoic acid, reactive oxygen species and reactive aldehyde metabolisms [63]. The expression and activity of these enzymes have been used, together with other biomarkers, for the identification of CSCs in different types of cancers. For instance, breast cancer stem cells are identified using specific antibodies for CD44 (cluster of differentiation 44) and CD24 (cluster of differentiation 24), and an ADELFLUOR assay to detect ALDH activity. CD44+/CD24/ALDH+ populations have been shown to possess characteristics of stem cells such as self-renewal, high proliferative potential, clonogenicity and multipotency [67,68,69,70]. In summary, ALDHs are associated with CSCs-mediated chemoresistance; however, their mechanisms of action are not well known. Thus, determining their modes of action will certainly help develop better targeting of these enzymes to prevent chemoresistance.

3.4. Resistance to Apoptosis

CCSs have been shown to resist apoptosis mediated by the intrinsic or mitochondria-dependent death pathway, and by the extrinsic cell death receptors pathway [71]. For instance, human glioma and leukaemia CSCs express lower levels of Fas and Fas ligand (Fas-L) resulting in resistance to extrinsic receptor-mediated cell death [72,73,74,75,76,77]. CSCs have also been shown to express higher levels of apoptotic inhibitors. The cellular Fas-associated death domain-like IL-1β-converting enzyme (FLICE)-inhibitory protein (c-FLIP) is highly expressed in leukaemia stem cells, and breast and glioblastoma CSCs [78,79,80]. C-FLIP averts apoptosis by preventing the formation of DISC (death-inducing signaling complex) and consequent activation of the caspase cascade via its interaction with FADD (Fas-associated death domain), caspase-8 or 10 and DR5 (Death receptor 5) [81,82]. Beside the extrinsic apoptotic pathway, the pro-survival protein Bcl-2 that is involved in the mitochondria-dependent death pathway, via its inhibitory interactions with the proapoptotic Bax (Bcl-2-associated x) and Bak (Bcl-2 antagonist killer), was found overexpressed in leukaemia, glioma and glioblastoma stem cells [83,84,85]. In summary, the inactivation of intrinsic and extrinsic cell death pathways ensures a selective survival advantage for CSCs.

3.5. DNA Repair Response

Another survival advantage in stressful conditions is related to CSCs’ capacity to timely activate the DNA damage sensor and repair machinery [86]. This process involves DNA repair pathways that include the nucleotide excision repair (NER), base excision repair (BER), mismatch repair (MMR), direct repair, and the double-strand break (DSB) recombinational repair. Evidence of CSCs’ resistance to radiotherapy was provided by a study that demonstrated the capacity of glioblastoma stem cells to efficiently activate the ATM serine/threonine kinase (ATM) and the DNA damage checkpoint protein kinase (Chk1) in response to ionising radiations [87]. Another study established the capacity of non-small cell lung cancer (NSCLC) stem cells to resist treatment with chemotherapeutic drugs by activating the Chk1 [88]. Mammosphere cells from the MCF-7 breast cancer cell line have been shown to have lower reactive oxygen species and a more active DNA single-strand break repair (SSBR) pathway, potentially associated to a higher level of expression of the key SSBR protein, human AP endonuclease 1 (Ape1) [89]. In conclusion, the enhanced DNA damage repair exhibited by CSCs appears to protect them from chemotherapy and radiotherapy-mediated apoptosis and adds another layer of protection to their survival capacities. Therefore, a better understanding of the mechanisms involved in DNA repair response will significantly contribute to improved therapy response and clinical outcome of patients treated with radiotherapy and DNA-targeting chemotherapies.

4. CSC Microenvironment

Stem cell niches represent areas of tissue that provide specific microenvironments, which maintain and promote CSCs’ capacity to self-renew and to generate differentiated progenies [90]. The concept of stem cell niche was first formulated by Schofield who demonstrated that successful transplants could only be successful if harvested from the bone marrow [91,92]. He also proposed that the stem cell niche is required for determining stem cell fate, as the behaviour of stem cells is influenced by their association with other cells within the niche. This concept was validated by other groups studying different tissues from invertebrate and vertebrate models [93,94]. This concept is applicable to cancer stem cells where the interaction with these niches is necessary for the maintenance of CSC populations. The CSC microenvironment is a highly heterogeneous complex comprised of cells such as stromal cells, immune cells and epithelial cells, and a network of extracellular macromolecules which provides support for the cells within the extracellular matrix (EC) (Figure 2) [95,96,97]. The cells found in the niche promote the growth, maintenance and differentiation of cancer stem cells. Currently, cancer therapies have been less successful as most of these drugs target the Bulk population of cancer cells, leaving cancer stem cell populations unaffected [98,99]. Therefore, understanding the relationship between CSCs and their microenvironment could help develop more efficient strategies to eliminate CSCs.

4.1. Cancer-Associated Fibroblasts (CAFs)

Although CAFs provide a mechanical supportive role for CSCs via enhanced production of fibrillary collagen, these cells also secrete the cytokine CXCL12 (C-X-C motif chemokine ligand 12) and growth factors such as the hepatocyte growth factor (HGF), the vascular endothelial growth factor (VEGF) and the platelet-derived growth factor (PDGF), which significantly contribute to CSCs’ increased proliferation, invasion and metastasis [97,100,101]. CAFs are also involved in cellular heterogenicity through secreted TGFβ1 (transforming growth factor beta 1), which promotes CSC related epithelial-mesenchymal transition (EMT), an early step of the invasive and metastatic process [102].

4.2. Immune Cells

Several immune cells contribute to the chronic inflammatory status of the CSC microenvironment, which enhances tumour proliferation, invasion and metastasis [103]. For instance, tumour-associated macrophages (TAMs) and myeloid-derived suppressor cells (MDSCs) secrete transforming growth factor beta (TGF-β), which contributes to EMT, invasion and metastasis [104,105,106,107]. In addition, immune cells within the microenvironment enhance tumour evasion via a variety of mechanisms. TAMs, via secreted TGFβ, recruit Tregs within the niche and contribute to immunosuppression. In a similar fashion, MDSCs, which secrete growth factors such as TGFβ, and cytokines recruit T helper 17 cells to promote their immunosuppressive activity [108]. Although we provide a very brief description here of the role of immune cells within the tumour microenvironment, a deeper discussion of their roles in this context is beyond the scope of the review.

4.3. Mesenchymal Stem Cells (MSCs)

MSCs are multipotent stromal cells that can differentiate into various cell types such as osteocytes, adipocytes and chondrocytes. These cells can migrate to chronic inflammatory sites such as cancer, where they contribute to metastasis by secreting TGFβ which promotes EMT [109]. An involvement in the colonisation process of secondary cancer sites has also been reported for metastatic breast, prostate and lung cancers [110,111,112,113]. MSCs have been shown to promote gastric cancer proliferation and angiogenesis via the secretion of VEGF, macrophage inflammatory protein-2 (MIP-2), TGF-β1, and the pro-inflammatory cytokines interleukin IL-6 and IL-8 [114]. Finally, MSCs may originate CAFs, which further contribute to CSC microenvironment cell heterogenicity and its metastatic potential [115,116,117].

4.4. Endothelial Cells

Angiogenesis is key for the CSC microenvironment as it delivers nutrients for CSC metabolism that are necessary for their self-renewal, and invasive and metastatic abilities. Through tumour vessels, immune cells such as T regulatory cells (Tregs) are recruited in situ and contribute to immune suppression [118]. Endothelial cells, together with perivascular cells, are the building blocks of vessels, and are stimulated by angiogenic factors such as the vascular endothelial growth factor (VEGF) and within a hypoxic environment such as cancer. This results in increased tumour vasculature and enhanced tumour growth and metastasis [119]. Additionally, tumour endothelial cells have been shown to secrete cytokines such as IL-3, granulocyte-CSF, granulocyte-macrophage-CSF (GM-CSF), IL-1, IL-6, VEGF-A, and bFGF, which promote and maintain CSC self-renewal and CSC-mediated progression [120,121].

5. Cancer Stemness, EMT and Tumour Progression

The main interest in investigating CSCs is related to their capacity to resist chemotherapy and radiotherapy, which results in cancer recurrence that is extremely difficult to overcome using currently available therapies. CSCs can re-initiate the tumorigenic process by generating a new tumour via their self-renewal and generation of differentiated cancer progenies. Another important property is their high invasive and metastatic potential, which involves a cellular process named EMT. Epithelial cells are connected to each other via lateral cell-cell junctions that include desmosomes, adherens junctions, tight junctions and gap junctions. During EMT, epithelial cells begin to trans-differentiate and start losing their lateral junctions and express molecules associated with the mesenchymal phenotype, resulting in the generation of mesenchymal cells [122]. These later play an essential role in the invasion process of adjacent tissues and distant organs via metastasis. The cytokine TGFβ, which is secreted in the CSC microenvironment by MSCs, CAFs and immune cells such as TAMs and MDSCs, plays an essential role in EMT-mediated CSC invasion [123]. TGF-β binding to its receptors results in the activation of the SMAD (small and mothers against decapentaplegic) proteins that translocate to the nucleus and activate target genes involved in EMT such as SNAIL, bHLH and ZEB factors [124,125]. These pro-EMT factors promote the expression of mesenchymal proteins such as N-cadherin, fibronectin and metalloproteinases (MMPs) [123]. Beside the TGFβ signalling, other signalling pathways, such as the WNT (wingless-related integration site), Notch (neurogenic locus notch homolog) and Hedgehog (HH) pathways, also participate in EMT. WNT signalling promotes EMT by stabilising β-catenin through the inhibition of glycogen synthase kinase-3β (GSK3β), resulting in β-catenin translocation to the nucleus and activation of the transcription factors lymphoid enhancer-binding factor 1 (LEF) and T cell factor (TCF), thus promoting a gene expression programme that favours EMT [126]. SHH signalling increases SNAIL1 expression in epithelial cancers leading to increased EMT and invasion [127]. Notch signalling may promote EMT through the Notch intracellular domain that directly activates SNAIL2 expression [128]. Finally, a crosstalk between TGFβ, WNT, Notch, Hedgehog (HH) and other signalling pathways such as RTK (receptor tyrosine kinase) signalling pathways may take place to promote EMT and invasion within the tumour microenvironment [129,130,131]. In summary, several pathways are involved in regulating EMT, and their targeting may require combination therapies that are directed against different components of this complex network, and with limited toxic effects.

6. CSC Metabolism

Cells require energy to perform their own processes of growth, division and survival. This need is met by the glycolysis metabolic pathway that converts glucose into pyruvate, two energy molecules of ATP and one of NADH (nicotinamide adenine dinucleotide hydrogen) [132]. The pyruvate is later transported in the mitochondria where it is converted into Acetyl-coA. The coenzyme produces the energy molecules NADH and FADH2 (flavin adenine dinucleotide hydrogen 2) through the Krebs cycle, which is also called the tricarboxylic acid cycle. The hydrogen ions that are used for the electron transport chain via the process of oxidative phosphorylation (OXPHOS) generate 36 molecules of ATPs. CSCs have been shown to have higher glycolytic metabolism when compared to normal and cancer cells. This difference could be due to higher expression of glycolytic enzymes such as GLUT1 (glucose transporter 1), HK-1 (Hexokinase-1) and PDK-1 (Pyruvate Dehydrogenase Kinase 1), and a higher consumption of glucose [132,133,134]. However, several studies challenged the concept of CSCs’ preferential reliance on the glycolytic metabolism pathway for energy production. These studies suggested that OXPHOS is more likely to be the main source of ATP due to increased mitochondrial ROS (reactive oxygen species), higher rates of oxygen consumption, and overall increase in mitochondrial function [135,136,137,138,139,140,141]. Overall, these opposing results could be explained by heterogenous differences in tissue origins, metabolic phenotypes and CSC adaptivity to environmental changes [142]. Therefore, until there is a better understanding of CSC metabolisms, developing targeted therapies remains a future therapeutic possibility.

7. Targeting Strategies of CSCs

CSCs are identified and isolated using cell markers whose expression create tissue-specific signatures. While some markers are commonly expressed by different types of cancer, others are tissue specific (Table 1). In this part, we provide an overview of the different approaches that were clinically employed to target CSCs and the results, so far, of past and ongoing clinical trials (Table 2).

7.1. Targeting Cell Surface Biomarkers of CSCs

7.1.1. CD133

CD133 (cluster of differentiation 133 or human prominin-1), a marker that was first identified on the cell surface of CD34+ hematopoietic stem cells [143], has also been used to isolate populations of CSCs from glioblastoma, prostate, pancreas, ovarian, colon, lung and liver cancer (Table 1). The therapeutic targeting potential of CD133 was evaluated for advanced cholangiosarcoma, in a phase 2 clinical trial (NCT02541370) (CCA), and using an immunotherapeutic approach based on CAR-T cells (chimeric antigen receptor T cells) that were transduced with the anti-CD133 (CD133 CAR-T) and anti-EGFR (epidermal growth factor receptor) (CART-EGFR). Although the authors suggested that CART cocktail immunotherapy (CD133 CART and CART-EGFR) may be feasible for the treatment of CCA and other solid malignancies, resolving limitations that are associated with toxicities would need further investigation (Table 2) [144].

7.1.2. CD44

CD44 is a cell surface adhesion receptor that is expressed by several types of cancers. The function of CD44 in growth and migration is dependent on its interaction with different ligands such as hyaluronic acid (HA), osteopontin (OPN), collagens and metalloproteases (MMPs) [145]. Targeting strategies for CD44 have led to the development of an antibody-based therapy that was clinically tested in a phase I study and on patients with refractory AML (acute myeloic leukeamia). The therapy using the humanised monoclonal antibody RG7356 resulted in limited clinical activity; however, the study validates the safety of the clinical use of the drug and suggests testing a combination therapy with other available drugs such as cytarabine (Table 2) [146].

7.1.3. EpCAM

The epithelial cell adhesion molecule (EpCAM) is a transmembrane glycoprotein that is expressed by most human carcinomas [147]. The first EpCAM monoclonal antibody (edrecolomab) was tested in patients with metastatic colorectal cancer or in adjuvant settings and had a limited response in a defined number of patients with metastatic disease [148,149]. Unfortunately, subsequent larger studies could not confirm edrecolomab’s clinical activity in the adjuvant setting [150,151]. Other antibodies that were developed such as 3622W94 and ING-1 were discontinued in clinical trials due to poor tolerability (Table 2) [152]. Adecatumumab, a recombinant human IgG1 monoclonal, that was clinically tested in patients with metastatic breast cancer, resulted in a more favourable response in patients with higher expression of EpCAM compared to patients who had lower expression on their cell surface (Table 2) [153].

7.1.4. CD47

The integrin associated protein (IAP) or CD47 (cluster of differentiation 47) was first identified as a β3 integrin interacting partner and was later found to interact with trombospondin-1 (TSP-1) and signal regulatory protein-alpha (SIRPα). Through these interactions, CD47 is involved in the regulation of various cellular functions such as cell migration, axon extension, cytokine production and T cell activation [154,155,156,157,158]. CD47 is overexpressed in most cancer types and was shown to promote cancer cell invasion and metastasis [159,160,161,162,163,164,165,166,167]. Two monoclonal antibodies have been developed to directly target CD47 in different types of cancer and were clinically tested as monotherapy or in combination with other monoclonal antibodies such as cetuximab and rituximab (Table 2) [168,169,170,171]. The first humanised anti-CD47 monoclonal antibody, Hu5F9-G4, was administered to patients with various solid tumours in a phase I clinical trial (NCT02216409). The trial aimed at optimising the dosage, the pharmacodynamic and the pharmacokinetic of the administered antibody. Although promising, the results were not statistically significant, and another clinical trial was performed with results not yet available (NCT02678338). The other monoclonal anti-CD47 monoclonal antibody CC-90002 was administered to patients with leukemia and myeloid and acute myelodysplastic syndromes in a phase I clinical trial, and the results are not yet published (NCT02641002). Both Hu5F9-G4 and CC-90002 are being clinically tested in combination therapies with the anti-EGFR antibody (cetuximab) or the anti-CD20 (cluster of differentiation 20) monoclonal antibody for the treatment of solid and haematological malignancies, and the trials are still ongoing (NCT02953782, NCT02953509, NCT02367196) (Table 2).

7.2. Dendritic Cell-Based Vaccines

Dendritic cells (DCs) are antigen presenting cells that initiate and maintain immune responses through their essential function in anti-cancer immunity [172]. DCs present cancer antigens to effector T-cells, which leads to the targeting and eradication of cancer cells. They can also stimulate immunological memories that control cancer recurrence. These properties favour use in the development of dendritic cell-based immunotherapies [172,173]. Several dendritic cell-based vaccines have been tested in patients with glioblastoma multiforme in phase I and II clinical trials (Table 2). A phase I clinical trial study of a dendritic cell vaccine for newly diagnosed or recurrent glioblastoma patients resulted in adverse events and treatment-related toxicities (NCT02010606). This study used patients own immune-stimulating dendritic cells, which were isolated and treated to promote an immune response against cancer stem cells and injected under the skin in a series of vaccinations. Another phase I study investigated the use of CD133 peptides for a dendritic cell-based vaccine (ICT-121 DC vaccine), which was clinically tested in patients with glioblastoma multiforme (NCT02049489). In this study CD133 peptides were used to load DCs that were injected in patients to elicit an immune response. So far, the results of this study have not been published. An ongoing phase II study investigates the efficacy of autologous dendritic cells (DCs) that were loaded with autogeneic glioma stem-like cells (A2B5+) and were administered as a vaccination in adults with glioblastoma multiforme (primary or secondary) (NCT01567202). Finally, EPCAM-targeted CAR-T cells have been generated using genetically modified T cells that express a chimeric antigen receptor (CAR), and that bind to tumour cells expressing the EPCAM protein on the cell surface [174]. These CAR-Ts are being clinically tested in patients and for different types of solid tumours (NCT02729493, NCT02725125, NCT02915445 and NCT03013712).

8. Conclusions

Although progress has been made in developing cancer therapies that target CSCs, the specificity of the targeted antigens remains a challenge. Most of the therapies are directed against biomarkers that are not specifically expressed by CSCs and that are also found in normal cells within the body. This has unfortunate effects on increasing toxicity in treated patients. For instance, therapies that target CD133 resulted in adverse events and treatment-related toxicities in glioblastoma patients who were treated with a CD133-specific dendritic vaccine (NCT02010606). However, other ongoing therapies such as the ones combining two monoclonal antibodies or using EPCAM as a target may lead to better therapies. Making the data of the clinical trials available to the scientific and clinical communities will certainly help in sharing information for the development of more efficient targeting therapies (NCT02678338, NCT02641002, NCT02729493, NCT02725125, NCT02915445 and NCT03013712). Although cancer bi-therapies are certainly a better choice, further studies are required to identify more specific targets with a much lower toxicity when tested in clinical trials. Large scale analysis of tumour samples from different patients and cohorts using genomics, proteomics and data mining approaches (Bioinformatics) could lead to the identification of CSC-specific genes or antigen signatures that will allow the development of more efficient CSC targeted therapies with lower associated toxicities for treated patients. Furthermore, these approaches could also provide patient specific signatures that will allow a personalised therapy and reduced cost and toxicity.
Another aspect that has not been discussed in this review, is the role of epigenetics in regulating the role of CSCs in cancer initiation and progression. Non-coding RNAs such as microRNAs (miRNAs) and long non-coding RNAs (lncRNAs) function by regulating the expression of several CSC-specific genes. Other epigenetic regulators involve histone acetyl transferases (HATs) and histone deacetylases, which also play important roles in maintaining CSC stemness and differentiation [175,176]. Therefore, developing therapies, such as gene targeting, nanoparticles-based therapies, and HDAC inhibitors could also constitute a potential way forward for targeting CSCs.
In summary, whether using drugs, immunotherapies or gene therapies, the specificity of the targets is key for a successful eradication of CSCs and cancer recurrence.

Funding

This research received no external funding.

Acknowledgments

We would like to thank the John and Lucille van Geest Foundation and IMMUTRAIN for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Global Health Observatory (GHO) data. Available online: https://www.who.int/gho/publications/world_health_statistics/2019/en/ (accessed on 01 July 2019).
  2. Regad, T. Tissue-specific cancer stem cells: Reality or a mirage? Transl. Med. Rep. 2017, 1. [Google Scholar] [CrossRef]
  3. Regad, T.; Sayers, T.J.; Rees, R.C. Principles of Stem Cell Biology and Cancer: Future Applications and Therapeutics; Wiley-Blackwell: Oxford, UK, 2015. [Google Scholar]
  4. Yu, Z.; Pestell, T.G.; Lisanti, M.P.; Pestell, R.G. Cancer stem cells. Int. J. Biochem. Cell Boil. 2012, 44, 2144–2151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Eyler, C.E.; Rich, J.N. Survival of the fittest: cancer stem cells in therapeutic resistance and angiogenesis. J. Clin. Oncol. 2008, 26, 2839–2845. [Google Scholar] [CrossRef] [PubMed]
  6. Baumann, M.; Krause, M.; Hill, R. Exploring the role of cancer stem cells in radioresistance. Nat. Rev. Cancer 2008, 8, 545–554. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, J. Cancer Stem Cells and Chemoresistance: The Smartest Survives the Raid. Pharmacol. Ther. 2016, 160, 145–158. [Google Scholar] [CrossRef] [PubMed]
  8. La Noce, M.; Paino, F.; Mele, L.; Papaccio, G.; Regad, T.; Lombardi, A.; Papaccio, F.; Desiderio, V.; Tirino, V. HDAC2 depletion promotes osteosarcoma’s stemness both in vitro and in vivo: A study on a putative new target for CSCs directed therapy. J. Exp. Clin. Cancer Res. 2018, 37, 296. [Google Scholar] [CrossRef] [PubMed]
  9. Takahashi, R.-U.; Miyazaki, H.; Ochiya, T. The role of microRNAs in the regulation of cancer stem cells. Front. Genet. 2014, 4, 295. [Google Scholar] [CrossRef] [Green Version]
  10. Oskarsson, T.; Batlle, E.; Massagué, J. Metastatic Stem Cells: Sources, Niches, and Vital Pathways. Cell Stem Cell 2014, 14, 306–321. [Google Scholar] [CrossRef] [Green Version]
  11. Voog, J.; Jones, D.L. Stem Cells and the Niche: A Dynamic Duo. Cell Stem Cell 2010, 6, 103–115. [Google Scholar] [CrossRef] [Green Version]
  12. Myron, S. Rudolf Virchow. Emerg Infect Dis. 2008, 14, 1480–1481. [Google Scholar] [Green Version]
  13. Welte, Y.; Adjaye, J.; Lehrach, H.R.; Regenbrecht, C.R. Cancer stem cells in solid tumors: Elusive or illusive? Cell Commun. Signal. 2010, 8, 6. [Google Scholar] [CrossRef] [PubMed]
  14. Bergsagel, D.E.; McCulloch, E.A.; Park, C.H. Mouse Myeloma Tumor Stem Cells: A Primary Cell Culture Assay2. J. Natl. Cancer Inst. 1971, 46, 411–422. [Google Scholar]
  15. Lapidot, T.; Sirard, C.; Vormoor, J.; Murdoch, B.; Hoang, T.; Caceres-Cortes, J.; Minden, M.; Paterson, B.; Caligiuri, M.A.; Dick, J.E. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 1994, 367, 645–648. [Google Scholar] [CrossRef] [PubMed]
  16. Al-Hajj, M.; Wicha, M.S.; Benito-Hernandez, A.; Morrison, S.J.; Clarke, M.F. Prospective identification of tumorigenic breast cancer cells. Proc. Natl. Acad. Sci. USA 2003, 100, 3983–3988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Schatton, T.; Murphy, G.F.; Frank, N.Y.; Yamaura, K.; Waaga-Gasser, A.M.; Gasser, M.; Zhan, Q.; Jordan, S.; Duncan, L.M.; Weishaupt, C.; et al. Identification of cells initiating human melanomas. Nature 2008, 451, 345–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Kumar, D.; Gorain, M.; Kundu, G.; Kundu, G.C. Therapeutic implications of cellular and molecular biology of cancer stem cells in melanoma. Mol. Cancer 2017, 16, 7. [Google Scholar] [CrossRef] [PubMed]
  19. Mathieu, M.G.; Miles, A.K.; Ahmad, M.; Buczek, M.E.; Pockley, A.G.; Rees, R.C.; Regad, T. The helicase HAGE prevents interferon-alpha-induced PML expression in ABCB5+ malignant melanoma-initiating cells by promoting the expression of SOCS1. Cell Death Disease 2014, 5, e1061. [Google Scholar] [CrossRef] [PubMed]
  20. Singh, S.K.; Clarke, I.D.; Terasaki, M.; Bonn, V.E.; Hawkins, C.; Squire, J.; Dirks, P.B. Identification of a cancer stem cell in human brain tumors. Cancer Res. 2003, 63, 5821–5828. [Google Scholar]
  21. Ricci-Vitiani, L.; Lombardi, D.G.; Pilozzi, E.; Biffoni, M.; Todaro, M.; Peschle, C.; De Maria, R. Identification and expansion of human colon-cancer-initiating cells. Nature 2007, 445, 111–115. [Google Scholar] [CrossRef]
  22. Dalerba, P.; Dylla, S.J.; Park, I.K.; Liu, R.; Wang, X.; Cho, R.W.; Hoey, T.; Gurney, A.; Huang, E.H.; Simeone, D.M.; et al. Phenotypic characterization of human colorectal cancer stem cells. Proc. Natl. Acad. Sci. USA 2007, 104, 10158–10163. [Google Scholar] [CrossRef] [Green Version]
  23. Yeung, T.M.; Gandhi, S.C.; Wilding, J.L.; Muschel, R.; Bodmer, W.F. Cancer stem cells from colorectal cancer-derived cell lines. Proc. Natl. Acad. Sci. USA 2010, 107, 3722–3727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Vermeulen, L.; Todaro, M.; de Sousa Mello, F.; Sprick, M.R.; Kemper, K.; Perez Alea, M.; Richel, D.J.; Stassi, G.; Medema, J.P. Single-cell cloning of colon cancer stem cells reveals a multi-lineage differentiation capacity. Proc. Natl. Acad. Sci. USA 2008, 105, 13427–13432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bekaii-Saab, T.; El-Rayes, B. Identifying and targeting cancer stem cells in the treatment of gastric cancer. Cancer 2017, 123, 1303–1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Singh, S.R. Gastric cancer stem cells: A novel therapeutic target. Cancer Lett. 2013, 338, 110–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Chinn, S.B.; Darr, O.A.; Owen, J.H.; Bellile, E.; McHugh, J.B.; Spector, M.E.; Papagerakis, S.M.; Chepeha, D.B.; Bradford, C.R.; Carey, T.E.; et al. Cancer stem cells: Mediators of tumorigenesis and metastasis in head and neck squamous cell carcinoma. Head Neck 2015, 37, 317–326. [Google Scholar] [CrossRef] [PubMed]
  28. Foster, R.; Buckanovich, R.J.; Rueda, B.R. Ovarian cancer stem cells: Working towards the root of stemness. Cancer Lett. 2013, 338, 147–157. [Google Scholar] [CrossRef]
  29. Li, C.; Heidt, D.G.; Dalerba, P.; Burant, C.; Zhang, L.; Adsay, V.; Wicha, M.; Clarke, M.F.; Simeone, D.M.; Lepelletier, Y.; et al. Identification of Pancreatic Cancer Stem Cells. Cancer Res. 2007, 67, 1030–1037. [Google Scholar] [CrossRef] [Green Version]
  30. Hou, Y.; Zou, Q.; Ge, R.; Shen, F.; Wang, Y. The critical role of CD133 (+) CD44(+/high) tumor cells in hematogenous metastasis of liver cancers. Cell Res. 2012, 22, 259–272. [Google Scholar] [CrossRef]
  31. Yang, Z.F.; Ho, D.W.; Ng, M.N.; Lau, C.K.; Yu, W.C.; Ngai, P.; Chu, P.W.; Lam, C.T.; Poon, R.T.; Fan, S.T. Significance of CD90+ cancer stem cells in human liver cancer. Cancer Cell 2008, 13, 153–166. [Google Scholar] [CrossRef]
  32. Kimura, O.; Takahashi, T.; Ishii, N.; Inoue, Y.; Ueno, Y.; Kogure, T.; Fukushima, K.; Shiina, M.; Yamagiwa, Y.; Kondo, Y.; et al. Characterization of the epithelial cell adhesion molecule (EpCAM)+ cell population in hepatocellular carcinoma cell lines. Cancer Sci. 2010, 101, 2145–2155. [Google Scholar] [CrossRef]
  33. Haraguchi, N.; Ishii, H.; Mimori, K.; Tanaka, F.; Ohkuma, M.; Kim, H.M.; Akita, H.; Takiuchi, D.; Hatano, H.; Nagano, H.; et al. CD13 is a therapeutic target in human liver cancer stem cells. J. Clin. Investig. 2010, 120, 3326–3339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Chen, X.; Li, Q.; Liu, X.; Liu, C.; Liu, R.; Rycaj, K.; Zhang, D.; Liu, B.; Jeter, C.; Calhoun-Davis, T.; et al. Defining a population of stem-like human prostate cancer cells that can generate and propagate castration-resistant prostate cancer (CRPC). Clin. Cancer Res. 2016, 22, 4505–4516. [Google Scholar] [CrossRef] [PubMed]
  35. MacDonagh, L.; Gray, S.G.; Breen, E.; Cuffe, S.; Finn, S.P.; O’Byrne, K.J.; Barr, M.P. Lung cancer stem cells: The root of resistance. Cancer Lett. 2016, 372, 147–156. [Google Scholar] [CrossRef] [PubMed]
  36. Shackleton, M.; Quintana, E.; Fearon, E.R.; Morrison, S.J. Heterogeneity in Cancer: Cancer Stem Cells versus Clonal Evolution. Cell 2009, 138, 822–829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Batlle, E.; Clevers, H. Cancer stem cells revisited. Nat. Med. 2017, 23, 1124–1134. [Google Scholar] [CrossRef] [PubMed]
  38. Rich, J.N. Cancer stem cells: Understanding tumor hierarchy and heterogeneity. Medicine 2016, 95, S2–S7. [Google Scholar] [CrossRef] [PubMed]
  39. Gottesman, M.M.; Lavi, O.; Hall, M.D.; Gillet, J.-P. Towards a Better Understanding of the Complexity of Cancer Drug Resistance. Annu. Rev. Pharmacol. Toxicol. 2016, 56, 85–102. [Google Scholar] [CrossRef]
  40. Hemmati, H.D.; Nakano, I.; Lazareff, J.A.; Masterman-Smith, M.; Geschwind, D.H.; Bronner-Fraser, M.; Kornblum, H.I. Cancerous stem cells can arise from pediatric brain tumors. Proc. Natl. Acad. Sci. USA 2003, 100, 15178–15183. [Google Scholar] [CrossRef] [Green Version]
  41. Fang, D.; Nguyen, T.K.; Leishear, K.; Finko, R.; Kulp, A.N.; Hotz, S.; Van Belle, P.A.; Xu, X.; Elder, D.E.; Herlyn, M.; et al. A Tumorigenic Subpopulation with Stem Cell Properties in Melanomas. Cancer Res. 2005, 65, 9328–9337. [Google Scholar] [CrossRef]
  42. Collins, A.T.; Berry, P.A.; Hyde, C.; Stower, M.J.; Maitland, N.J. Prospective Identification of Tumorigenic Prostate Cancer Stem Cells. Cancer Res. 2005, 65, 10946–10951. [Google Scholar] [CrossRef] [Green Version]
  43. Bapat, S.A.; Mali, A.M.; Koppikar, C.B.; Kurrey, N.K. Stem and progenitor-like cells contribute to the aggressive behavior of human epithelial ovarian cancer. Cancer Res. 2005, 65, 3025–3029. [Google Scholar] [CrossRef] [PubMed]
  44. Saikawa. Tumor initiating potential of side population cells in human gastric cancer. Int. J. Oncol. 2009, 34, 1201–1207. [Google Scholar] [CrossRef] [Green Version]
  45. Ho, M.M.; Ng, A.V.; Lam, S.; Hung, J.Y. Side Population in Human Lung Cancer Cell Lines and Tumors Is Enriched with Stem-like Cancer Cells. Cancer Res. 2007, 67, 4827–4833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hermann, P.C.; Huber, S.L.; Herrler, T.; Aicher, A.; Ellwart, J.W.; Guba, M.; Bruns, C.J.; Heeschen, C. Distinct Populations of Cancer Stem Cells Determine Tumor Growth and Metastatic Activity in Human Pancreatic Cancer. Cell Stem Cell 2007, 1, 313–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Curtis, S.J.; Sinkevicius, K.W.; Li, D.; Lau, A.N.; Roach, R.R.; Zamponi, R.; Woolfenden, A.E.; Kirsch, D.G.; Wong, K.-K.; Kim, C.F. Primary tumor genotype is an important determinant in identification of lung cancer propagating cells. Cell Stem Cell 2010, 7, 127–133. [Google Scholar] [CrossRef] [PubMed]
  48. Clarkson, B.; Strife, A.; Fried, J.; Sakai, Y.; Ota, K.; Ohkita, T.; Masuda, R. Studies of cellular proliferation in human leukemia. IV. Behavior of normal hematopoietic cells in 3 adults with acute leukemia given continuous infusions of3 H-thymidine for 8 or 10 days. Cancer 1970, 26, 1–19. [Google Scholar] [CrossRef]
  49. Moore, N.; Houghton, J.; Lyle, S. Slow-Cycling Therapy-Resistant Cancer Cells. Stem Cells Dev. 2012, 21, 1822–1830. [Google Scholar] [CrossRef] [Green Version]
  50. Vidal, S.J.; Rodriguez-Bravo, V.; Galsky, M.; Cordon-Cardo, C.; Domingo-Domenech, J. Targeting cancer stem cells to suppress acquired chemotherapy resistance. Oncogene 2013, 33, 4451–4463. [Google Scholar] [CrossRef] [Green Version]
  51. Ajani, J.A.; Song, S.; Hochster, H.S.; Steinberg, I.B. Cancer Stem Cells: The Promise and the Potential. Semin. Oncol. 2015, 42, S3–S17. [Google Scholar] [CrossRef]
  52. Abdullah, L.N.; Chow, E.K.-H. Mechanisms of chemoresistance in cancer stem cells. Clin. Transl. Med. 2013, 2, 3. [Google Scholar] [CrossRef]
  53. Begicevic, R.-R.; Falasca, M. ABC Transporters in Cancer Stem Cells: Beyond Chemoresistance. Int. J. Mol. Sci. 2017, 18, 2362. [Google Scholar] [CrossRef]
  54. Kolenda, J.; Jensen, S.S.; Aaberg-Jessen, C.; Christensen, K.; Andersen, C.; Brünner, N.; Kristensen, B.W. Effects of hypoxia on expression of a panel of stem cell and chemoresistance markers in glioblastoma-derived spheroids. J. Neuro-Oncol. 2010, 103, 43–58. [Google Scholar] [CrossRef]
  55. Deeley, R.G.; Cole, S.P. Substrate recognition and transport by multidrug resistance protein 1 (ABCC1). FEBS Lett. 2005, 580, 1103–1111. [Google Scholar] [CrossRef] [Green Version]
  56. Zhou, S.-F.; Wang, L.-L.; Di, Y.M.; Xue, C.C.; Duan, W.; Li, C.G.; Li, Y. Substrates and inhibitors of human multidrug resistance associated proteins and the implications in drug development. Curr. Med. Chem. 2008, 15, 1981–2039. [Google Scholar] [CrossRef]
  57. Shaffer, B.C.; Gillet, J.-P.; Patel, C.; Baer, M.R.; Bates, S.E.; Gottesman, M.M. Drug resistance: Still a daunting challenge to the successful treatment of AML. Drug Resist. Updat. 2012, 15, 62–69. [Google Scholar] [CrossRef] [Green Version]
  58. Xi, G.; Hayes, E.; Lewis, R.; Ichi, S.; Mania-Farnell, B.; Shim, K.; Takao, T.; Allender, E.; Mayanil, C.S.; Tomita, T. CD133 and DNA-PK regulate MDR1 via the PI3K- or Akt-NF-κB pathway in multidrug-resistant glioblastoma cells in vitro. Oncogene 2015, 35, 241–250. [Google Scholar] [CrossRef]
  59. Huang, B.; Fu, S.J.; Fan, W.Z.; Wang, Z.H.; Chen, Z.B.; Guo, S.J.; Chen, J.X.; Qiu, S.P. PKCε inhibits isolation and stemness of side population cells via the suppression of ABCB1 transporter and PI3K/Akt, MAPK/ERK signaling in renal cell carcinoma cell line 769P. Cancer Lett. 2016, 376, 148–154. [Google Scholar] [CrossRef]
  60. Johnson, R.A.; Shepard, E.M.; Scotto, K.W. Differential Regulation of MDR1 Transcription by the p53 Family Members. J. Boil. Chem. 2005, 280, 13213–13219. [Google Scholar] [CrossRef] [Green Version]
  61. Chou, J.-L.; Huang, R.-L.; Shay, J.; Chen, L.-Y.; Lin, S.-J.; Yan, P.S.; Chao, W.-T.; Lai, Y.-H.; Lai, Y.-L.; Chao, T.-K.; et al. Hypermethylation of the TGF-β target, ABCA1 is associated with poor prognosis in ovarian cancer patients. Clin. Epigenet. 2015, 7, 1. [Google Scholar] [CrossRef]
  62. Sládek, N.E. Human aldehyde dehydrogenases: Potential pathological, pharmacological, and toxicological impact. J. Biochem. Mol. Toxicol. 2003, 17, 7–23. [Google Scholar] [CrossRef]
  63. Ikawa, M.; Impraim, C.C.; Wang, G.; Yoshida, A. Isolation and characterization of aldehyde dehydrogenase isozymes from usual and atypical human livers. J. Boil. Chem. 1983, 258, 6282–6287. [Google Scholar]
  64. Ginestier, C.; Hur, M.H.; Charafe-Jauffret, E.; Monville, F.; Dutcher, J.; Brown, M.; Jacquemier, J.; Viens, P.; Kleer, C.G.; Liu, S.; et al. ALDH1 is a marker of normal and malignant human mammary stem cells and a predictor of poor clinical outcome. Cell Stem Cell 2007, 1, 555–567. [Google Scholar] [CrossRef]
  65. Fitzgerald, T.L.; Rangan, S.; Dobbs, L.; Starr, S.; Sigounas, G. The impact of Aldehyde dehydrogenase 1 expression on prognosis for metastatic colon cancer. J. Surg. Res. 2014, 192, 82–89. [Google Scholar] [CrossRef]
  66. Vogler, T.; Kriegl, L.; Horst, D.; Engel, J.; Sagebiel, S.; Schäffauer, A.J.; Kirchner, T.; Jung, A. The expression pattern of aldehyde dehydrogenase 1 (ALDH1) is an independent prognostic marker for low survival in colorectal tumors. Exp. Mol. Pathol. 2012, 92, 111–117. [Google Scholar] [CrossRef]
  67. Deng, L.; Zhou, J.-F.; Sellers, R.S.; Li, J.-F.; Nguyen, A.V.; Wang, Y.; Orlofsky, A.; Liu, Q.; Hume, D.A.; Pollard, J.W.; et al. A Novel Mouse Model of Inflammatory Bowel Disease Links Mammalian Target of Rapamycin-Dependent Hyperproliferation of Colonic Epithelium to Inflammation-Associated Tumorigenesis. Am. J. Pathol. 2010, 176, 952–967. [Google Scholar] [CrossRef]
  68. Hessman, C.J.; Bubbers, E.J.; Billingsley, K.G.; Herzig, D.O.; Wong, M.H. Loss of expression of the cancer stem cell marker aldehyde dehydrogenase1 correlates with advanced stage colorectal cancer. Am. J. Surg. 2012, 203, 649–653. [Google Scholar] [CrossRef]
  69. Kim, M.P.; Fleming, J.B.; Wang, H.; Abbruzzese, J.L.; Choi, W.; Kopetz, S.; McConkey, D.J.; Evans, D.B.; Gallick, G.E. ALDH Activity Selectively Defines an Enhanced Tumor-Initiating Cell Population Relative to CD133 Expression in Human Pancreatic Adenocarcinoma. PLoS ONE 2011, 6, e20636. [Google Scholar] [CrossRef]
  70. Li, T.; Su, Y.; Mei, Y.; Leng, Q.; Leng, B.; Liu, Z.; Stass, S.A.; Jiang, F. ALDH1A1 is a marker for malignant prostate stem cells and predictor of prostate cancer patients’ outcome. Lab. Investig. 2009, 90, 234–244. [Google Scholar] [CrossRef]
  71. Wang, Y.-H.; Scadden, D.T. Harnessing the apoptotic programs in cancer stem-like cells. EMBO Rep. 2015, 16, 1084–1098. [Google Scholar] [CrossRef]
  72. Costello, R.T.; Mallet, F.; Chambost, H.; Sainty, D.; Arnoulet, C.; Gastaut, J.A.; Olive, D. The immunophenotype of minimally differentiated acute myeloid leukemia (AML-M0): Reduced immunogenicity and high frequency of CD34+/CD38− leukemic progenitors. Leukemia 1999, 13, 1513–1518. [Google Scholar] [CrossRef]
  73. Weller, M.; Frei, K.; Groscurth, P.; Krammer, P.H.; Yonekawa, Y.; Fontana, A. Anti-Fas/APO-1 antibody-mediated apoptosis of cultured human glioma cells. Induction and modulation of sensitivity by cytokines. J. Clin. Investig. 1994, 94, 954–964. [Google Scholar] [CrossRef]
  74. Frei, K.; Ambar, B.; Adachi, N.; Yonekawa, Y.; Fontana, A. Ex vivo malignant glioma cells are sensitive to Fas (CD95/APO-1) ligand-mediated apoptosis. J. Neuroimmunol. 1998, 87, 105–113. [Google Scholar] [CrossRef]
  75. Tao, J.; Qiu, B.; Zhang, D.; Wang, Y. Expression levels of Fas/Fas-L mRNA in human brain glioma stem cells. Mol. Med. Rep. 2012, 5, 1202–1206. [Google Scholar] [CrossRef]
  76. Eisele, G.; Roth, P.; Hasenbach, K.; Aulwurm, S.; Wolpert, F.; Tabatabai, G.; Wick, W.; Weller, M. APO010, a synthetic hexameric CD95 ligand, induces human glioma cell death in vitro and in vivo. Neuro-Oncology 2010, 13, 155–164. [Google Scholar] [CrossRef]
  77. Eisele, G.; Wolpert, F.; Decrey, G.; Weller, M. APO010, a synthetic hexameric CD95 ligand, induces death of human glioblastoma stem-like cells. Anticancer Res. 2013, 33, 3563–3571. [Google Scholar]
  78. Zobalova, R.; McDermott, L.; Stantic, M.; Prokopova, K.; Dong, L.-F.; Neuzil, J. CD133-positive cells are resistant to TRAIL due to up-regulation of FLIP. Biochem. Biophys. Res. Commun. 2008, 373, 567–571. [Google Scholar] [CrossRef]
  79. Dong, L.-F.; Freeman, R.; Liu, J.; Zobalova, R.; Marin-Hernandez, A.; Stantic, M.; Rohlena, J.; Valis, K.; Rodriguez-Enriquez, S.; Butcher, B.; et al. Suppression of Tumor Growth In vivo by the Mitocan α-tocopheryl Succinate Requires Respiratory Complex II. Clin. Cancer Res. 2009, 15, 1593–1600. [Google Scholar] [CrossRef]
  80. Ding, L.; Yuan, C.; Wei, F.; Wang, G.; Zhang, J.; Bellail, A.C.; Zhang, Z.; Olson, J.J.; Hao, C. Cisplatin Restores TRAIL apoptotic pathway in Glioblastoma-Derived Stem Cells through Up-regulation of DR5 and Down-regulation of c-FLIP. Cancer Investig. 2011, 29, 511–520. [Google Scholar] [CrossRef]
  81. Dutton, A.; Young, L.S.; Murray, P.G. The role of cellular flice inhibitory protein (c-FLIP) in the pathogenesis and treatment of cancer. Expert Opin. Ther. Targets 2006, 10, 27–35. [Google Scholar] [CrossRef]
  82. Shirley, S.; Micheau, O. Targeting c-FLIP in cancer. Cancer Lett. 2013, 332, 141–150. [Google Scholar] [CrossRef] [Green Version]
  83. Lauria, F.; Raspadori, D.; Rondelli, D.; Ventura, M.; Fiacchini, M.; Visani, G.; Forconi, F.; Tura, S. High bcl-2 expression in acute myeloid leukemia cells correlates with CD34 positivity and complete remission rate. Leukemia 1997, 11, 2075–2078. [Google Scholar] [CrossRef] [Green Version]
  84. Campos, L.; Oriol, P.; Sabido, O.; Guyotat, D. Simultaneous Expression of P-Glycoprotein and BCL-2 in Acute Myeloid Leukemia Blast Cells. Leuk. Lymphoma 1997, 27, 119–125. [Google Scholar] [CrossRef]
  85. Tóthová, E.; Kafková, A.; Stecová, N.; Fricová, M.; Guman, T.; Svorcová, E. Immune-mediated complications during interferon alpha therapy in chronic myelogenous leukemia. Neoplasma 2002, 49, 91–94. [Google Scholar]
  86. Maugeri-Saccà, M.; Bartucci, M.; De Maria, R. DNA Damage Repair Pathways in Cancer Stem Cells. Mol. Cancer Ther. 2012, 11, 1627–1636. [Google Scholar] [CrossRef]
  87. Bao, S.; Wu, Q.; McLendon, R.E.; Hao, Y.; Shi, Q.; Hjelmeland, A.B.; Dewhirst, M.W.; Bigner, D.D.; Rich, J.N. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef]
  88. Bartucci, M.; Svensson, S.; Romania, P.; Dattilo, R.; Patrizii, M.; Signore, M.; Navarra, S.; Lotti, F.; Biffoni, M.; Pilozzi, E.; et al. Therapeutic targeting of Chk1 in NSCLC stem cells during chemotherapy. Cell Death Differ. 2011, 19, 768–778. [Google Scholar] [CrossRef] [Green Version]
  89. Karimi-Busheri, F.; Rasouli-Nia, A.; Mackey, J.R.; Weinfeld, M. Senescence evasion by MCF-7 human breast tumor-initiating cells. Breast Cancer Res. 2010, 12, R31. [Google Scholar] [CrossRef]
  90. Morrison, S.J.; Spradling, A.C. Stem cells and niches: mechanisms that promote stem cell maintenance throughout life. Cell 2008, 132, 598–611. [Google Scholar] [CrossRef]
  91. Schofield, R. The relationship between the spleen colony-forming cell and the haemopoietic stem cell. Blood Cells 1978, 4, 7–25. [Google Scholar]
  92. Rashidi, N.M.; Lo Celso, C. Flying back to the nest: Intravital microscopy reveals how the niche can induce stemness. Intravital 2014, 3, e29653. [Google Scholar] [CrossRef]
  93. Ferraro, F.; Celso, C.L.; Scadden, D. Adult stem cells and their niches. Single Mol. Single Cell Seq. 2010, 695, 155–168. [Google Scholar] [CrossRef]
  94. Wu, T.; Dai, Y. Tumor microenvironment and therapeutic response. Cancer Lett. 2017, 387, 61–68. [Google Scholar] [CrossRef]
  95. Balkwill, F.; Capasso, M.; Hagemann, T. The tumor microenvironment at a glance. J. Cell Sci. 2012, 125, 5591–5596. [Google Scholar] [CrossRef] [Green Version]
  96. Korkaya, H.; Liu, S.; Wicha, M.S. Breast cancer stem cells, cytokine networks, and the tumor microenvironment. J. Clin. Investig. 2011, 121, 3804–3809. [Google Scholar] [CrossRef]
  97. Plaks, V.; Kong, N.; Werb, Z. The Cancer Stem Cell Niche: How Essential is the Niche in Regulating Stemness of Tumor Cells? Cell Stem Cell 2015, 16, 225–238. [Google Scholar] [CrossRef]
  98. Longley, D.; Johnston, P. Molecular mechanisms of drug resistance. J. Pathol. 2005, 205, 275–292. [Google Scholar] [CrossRef]
  99. Prieto-Vila, M.; Takahashi, R.-U.; Usuba, W.; Kohama, I.; Ochiya, T. Drug Resistance Driven by Cancer Stem Cells and Their Niche. Int. J. Mol. Sci. 2017, 18, 2574. [Google Scholar] [CrossRef]
  100. Junttila, M.R.; De Sauvage, F.J. Influence of tumour micro-environment heterogeneity on therapeutic response. Nature 2013, 501, 346–354. [Google Scholar] [CrossRef]
  101. Najafi, M.; Farhood, B.; Mortezaee, K. Cancer stem cells (CSCs) in cancer progression and therapy. J. Cell. Physiol. 2018, 234, 8381–8395. [Google Scholar] [CrossRef]
  102. Zhuang, J.; Lu, Q.; Shen, B.; Huang, X.; Shen, L.; Zheng, X.; Huang, R.; Yan, J.; Guo, H. TGFβ1 secreted by cancer-associated fibroblasts induces epithelial-mesenchymal transition of bladder cancer cells through lncRNA-ZEB2NAT. Sci. Rep. 2015, 5, 11924. [Google Scholar] [CrossRef]
  103. Cabarcas, S.M.; Mathews, L.A.; Farrar, W.L. The cancer stem cell niche-there goes the neighborhood? Int. J. Cancer 2011, 129, 2315–2327. [Google Scholar] [CrossRef]
  104. Buczek, M.E.; Miles, A.K.; Green, W.; Johnson, C.; Boocock, D.J.; Pockley, A.G.; Rees, R.C.; Hulman, G.; Van Schalkwyk, G.; Parkinson, R.; et al. Cytoplasmic PML promotes TGF-β-associated epithelial–mesenchymal transition and invasion in prostate cancer. Oncogene 2015, 35, 3465–3475. [Google Scholar] [CrossRef]
  105. Katsuno, Y.; Lamouille, S.; Derynck, R. TGF-β signaling and epithelial–mesenchymal transition in cancer progression. Curr. Opin. Oncol. 2013, 25, 76–84. [Google Scholar] [CrossRef]
  106. Kalluri, R.; Zeisberg, M. Fibroblasts in cancer. Nat. Rev. Cancer 2006, 6, 392–401. [Google Scholar] [CrossRef]
  107. Kalluri, R.; Weinberg, R.A. The basics of epithelial-mesenchymal transition. J. Clin. Investig. 2009, 119, 1420–1428. [Google Scholar] [CrossRef] [Green Version]
  108. Kitamura, T.; Qian, B.-Z.; Pollard, J.W. Immune cell promotion of metastasis. Nat. Rev. Immunol. 2015, 15, 73–86. [Google Scholar] [CrossRef]
  109. Ridge, S.M.; Sullivan, F.J.; Glynn, S.A. Mesenchymal stem cells: Key players in cancer progression. Mol. Cancer 2017, 16, 31. [Google Scholar] [CrossRef]
  110. Karnoub, A.E.; Dash, A.B.; Vo, A.P.; Sullivan, A.; Brooks, M.W.; Bell, G.W.; Richardson, A.L.; Polyak, K.; Tubo, R.; Weinberg, R.A. Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 2007, 449, 557–563. [Google Scholar] [CrossRef]
  111. Corcoran, K.E.; Trzaska, K.A.; Fernandes, H.; Bryan, M.; Taborga, M.; Srinivas, V.; Packman, K.; Patel, P.S.; Rameshwar, P. Mesenchymal Stem Cells in Early Entry of Breast Cancer into Bone Marrow. PLoS ONE 2008, 3, e2563. [Google Scholar] [CrossRef]
  112. Nabha, S.M.; Dos Santos, E.B.; Yamamoto, H.A.; Belizi, A.; Dong, Z.; Meng, H.; Saliganan, A.; Sabbota, A.; Bonfil, R.D.; Cher, M.L. Bone marrow stromal cells enhance prostate cancer cell invasion through type I collagen in an MMP-12 dependent manner. Int. J. Cancer 2008, 122, 2482–2490. [Google Scholar] [CrossRef] [Green Version]
  113. Duda, D.G.; Duyverman, A.M.M.J.; Kohno, M.; Snuderl, M.; Steller, E.J.A.; Fukumura, D.; Jain, R.K. Malignant cells facilitate lung metastasis by bringing their own soil. Proc. Natl. Acad. Sci. USA 2010, 107, 21677–21682. [Google Scholar] [CrossRef] [Green Version]
  114. Li, W.; Zhou, Y.; Yang, J.; Zhang, X.; Zhang, H.; Zhang, T.; Zhao, S.; Zheng, P.; Huo, J.; Wu, H. Gastric cancer-derived mesenchymal stem cells prompt gastric cancer progression through secretion of interleukin-8. J. Exp. Clin. Cancer Res. 2015, 34, 52. [Google Scholar] [CrossRef] [Green Version]
  115. Spaeth, E.L.; Dembinski, J.L.; Sasser, A.K.; Watson, K.; Klopp, A.; Hall, B.; Andreeff, M.; Marini, F. Mesenchymal Stem Cell Transition to Tumor-Associated Fibroblasts Contributes to Fibrovascular Network Expansion and Tumor Progression. PLoS ONE 2009, 4, e4992. [Google Scholar] [CrossRef]
  116. Mishra, P.J.; Mishra, P.J.; Humeniuk, R.; Medina, D.J.; Alexe, G.; Mesirov, J.P.; Ganesan, S.; Glod, J.W.; Banerjee, D. Carcinoma Associated Fibroblast Like Differentiation of Human Mesenchymal Stem Cells. Cancer Res. 2008, 68, 4331–4339. [Google Scholar] [CrossRef]
  117. Peng, Y.; Li, Z.; Li, Z. GRP78 secreted by tumor cells stimulates differentiation of bone marrow mesenchymal stem cells to cancer-associated fibroblasts. Biochem. Biophys. Res. Commun. 2013, 440, 558–563. [Google Scholar] [CrossRef]
  118. Hida, K.; Maishi, N.; Annan, D.A.; Hida, Y. Contribution of Tumor Endothelial Cells in Cancer Progression. Int. J. Mol. Sci. 2018, 19, 1272. [Google Scholar] [CrossRef]
  119. Fessler, E.; Borovski, T.; Medema, J.P. Endothelial cells induce cancer stem cell features in differentiated glioblastoma cells via bFGF. Mol. Cancer 2015, 14, 157. [Google Scholar] [CrossRef]
  120. Pirtskhalaishvili, G.; Nelson, J.B. Endothelium-derived factors as paracrine mediators of prostate cancer progression. Prostate 2000, 44, 77–87. [Google Scholar] [CrossRef]
  121. Butler, J.M.; Kobayashi, H.; Rafii, S. Instructive role of the vascular niche in promoting tumour growth and tissue repair by angiocrine factors. Nat. Rev. Cancer 2010, 10, 138–146. [Google Scholar] [CrossRef] [Green Version]
  122. Grünert, S.; Jechlinger, M.; Beug, H.; Gruenert, S. Diverse cellular and molecular mechanisms contribute to epithelial plasticity and metastasis. Nat. Rev. Mol. Cell Boil. 2003, 4, 657–665. [Google Scholar] [CrossRef]
  123. Lamouille, S.; Xu, J.; Derynck, R. Molecular mechanisms of epithelial–mesenchymal transition. Nat. Rev. Mol. Cell Boil. 2014, 15, 178–196. [Google Scholar] [CrossRef]
  124. Massague, J. TGFbeta signalling in context. Nature reviews. Mol. Cell Boil. 2012, 13, 616–630. [Google Scholar] [CrossRef]
  125. Thiery, J.P. Epithelial–mesenchymal transitions in tumour progression. Nat. Rev. Cancer 2002, 2, 442–454. [Google Scholar] [CrossRef]
  126. Niehrs, C. The complex world of WNT receptor signalling. Nat. Rev. Mol. Cell Boil. 2012, 13, 767–779. [Google Scholar] [CrossRef]
  127. Fendrich, V.; Waldmann, J.; Esni, F.; Ramaswamy, A.; Mullendore, M.; Buchholz, M.; Maitra, A.; Feldmann, G. Snail and Sonic Hedgehog activation in neuroendocrine tumors of the ileum. Endoc.-Relat. Cancer 2007, 14, 865–874. [Google Scholar] [CrossRef] [Green Version]
  128. Xie, M.; Zhang, L.; He, C.-S.; Xu, F.; Liu, J.-L.; Hu, Z.-H.; Zhao, L.-P.; Tian, Y. Activation of Notch-1 enhances epithelial-mesenchymal transition in gefitinib-acquired resistant lung cancer cells. J. Cell. Biochem. 2012, 113, 150–1513. [Google Scholar] [CrossRef]
  129. Shirakihara, T.; Horiguchi, K.; Miyazawa, K.; Ehata, S.; Shibata, T.; Morita, I.; Miyazono, K.; Saitoh, M. TGF-β regulates isoform switching of FGF receptors and epithelial–mesenchymal transition. EMBO J. 2011, 30, 783–795. [Google Scholar] [CrossRef]
  130. Uttamsingh, S.; Bao, X.; Nguyen, K.T.; Bhanot, M.; Gong, J.; Chan, J.L.-K.; Liu, F.; Chu, T.T.; Wang, L.-H. Synergistic effect between EGF and TGF-β1 in inducing oncogenic properties of intestinal epithelial cells. Oncogene 2007, 27, 2626–2634. [Google Scholar] [CrossRef]
  131. Wendt, M.K.; Smith, J.A.; Schiemann, W.P. Transforming Growth Factor-β-Induced Epithelial-Mesenchymal Transition Facilitates Epidermal Growth Factor-Dependent Breast Cancer Progression. Oncogene 2010, 29, 6485–6498. [Google Scholar] [CrossRef]
  132. Snyder, V.; Reed-Newman, T.C.; Arnold, L.; Thomas, S.M.; Anant, S. Cancer Stem Cell Metabolism and Potential Therapeutic Targets. Front. Oncol. 2018, 8, 203. [Google Scholar] [CrossRef]
  133. Emmink, B.L.; Verheem, A.; Van Houdt, W.J.; Steller, E.J.; Govaert, K.M.; Pham, T.V.; Piersma, S.R.; Rinkes, I.H.B.; Jimenez, C.R.; Kranenburg, O. The secretome of colon cancer stem cells contains drug-metabolizing enzymes. J. Proteom. 2013, 91, 84–96. [Google Scholar] [CrossRef]
  134. Hammoudi, N.; Ahmed, K.B.R.; Garcia-Prieto, C.; Huang, P. Metabolic alterations in cancer cells and therapeutic implications. Chin. J. Cancer 2011, 30, 508–525. [Google Scholar] [CrossRef] [Green Version]
  135. Ye, X.-Q.; Li, Q.; Wang, G.-H.; Sun, F.-F.; Huang, G.-J.; Bian, X.-W.; Yu, S.-C.; Qian, G.-S. Mitochondrial and energy metabolism-related properties as novel indicators of lung cancer stem cells. Int. J. Cancer 2011, 129, 820–831. [Google Scholar] [CrossRef]
  136. Janiszewska, M.; Suvà, M.L.; Riggi, N.; Houtkooper, R.H.; Auwerx, J.; Clément-Schatlo, V.; Radovanovic, I.; Rheinbay, E.; Provero, P.; Stamenkovic, I. Imp2 controls oxidative phosphorylation and is crucial for preserving glioblastoma cancer stem cells. Genes Dev. 2012, 26, 1926–1944. [Google Scholar] [CrossRef] [Green Version]
  137. Sancho, P.; Burgos-Ramos, E.; Tavera, A.; Kheir, T.B.; Jagust, P.; Schoenhals, M.; Barneda, D.; Sellers, K.; Campos-Olivas, R.; Graña, O.; et al. MYC/PGC-1α Balance Determines the Metabolic Phenotype and Plasticity of Pancreatic Cancer Stem Cells. Cell Metab. 2015, 22, 590–605. [Google Scholar] [CrossRef]
  138. Lagadinou, E.D.; Sach, A.; Callahan, K.; Rossi, R.M.; Neering, S.J.; Minhajuddin, M.; Ashton, J.M.; Pei, S.; Grose, V.; O’Dwyer, K.M.; et al. BCL-2 inhibition targets oxidative phosphorylation and selectively eradicates quiescent human leukemia stem cells. Cell Stem Cell 2013, 12, 329–341. [Google Scholar] [CrossRef]
  139. Pastò, A.; Bellio, C.; Pilotto, G.; Ciminale, V.; Silic-Benussi, M.; Guzzo, G.; Rasola, A.; Frasson, C.; Nardo, G.; Zulato, E.; et al. Cancer stem cells from epithelial ovarian cancer patients privilege oxidative phosphorylation, and resist glucose deprivation. Oncotarget 2014, 5, 4305–4319. [Google Scholar] [CrossRef] [Green Version]
  140. De Luca, A.; Fiorillo, M.; Peiris-Pagés, M.; Ozsvari, B.; Smith, D.L.; Sanchez-Alvarez, R.; Martinez-Outschoorn, U.E.; Cappello, A.R.; Pezzi, V.; Lisanti, M.P.; et al. Mitochondrial biogenesis is required for the anchorage-independent survival and propagation of stem-like cancer cells. Oncotarget 2015, 6, 14777–14795. [Google Scholar] [CrossRef]
  141. Vlashi, E.; Lagadec, C.; Vergnes, L.; Reue, K.; Frohnen, P.; Chan, M.; Alhiyari, Y.; Dratver, M.B.; Pajonk, F. Metabolic differences in breast cancer stem cells and differentiated progeny. Breast Cancer Res. Treat. 2014, 146, 525–534. [Google Scholar] [CrossRef] [Green Version]
  142. Zheng, J. Energy metabolism of cancer: Glycolysis versus oxidative phosphorylation (Review). Oncol. Lett. 2012, 4, 1151–1157. [Google Scholar] [CrossRef]
  143. Yin, A.H.; Miraglia, S.; Zanjani, E.D.; Almeida-Porada, G.; Ogawa, M.; Leary, A.G.; Olweus, J.; Kearney, J.; Buck, D.W. AC133, a novel marker for human hematopoietic stem and progenitor cells. Blood 1997, 90, 5002–5012. [Google Scholar]
  144. Feng, K.-C.; Guo, Y.-L.; Liu, Y.; Dai, H.-R.; Wang, Y.; Lv, H.-Y.; Huang, J.-H.; Yang, Q.-M.; Han, W.-D. Cocktail treatment with EGFR-specific and CD133-specific chimeric antigen receptor-modified T cells in a patient with advanced cholangiocarcinoma. J. Hematol. Oncol. 2017, 10, 4. [Google Scholar] [CrossRef]
  145. Senbanjo, L.T.; Chellaiah, M.A. CD44: A multifunctional cell surface adhesion receptor is a regulator of progression and metastasis of cancer cells. Front. Cell Dev. Boil. 2017, 5, 18. [Google Scholar] [CrossRef]
  146. Vey, N.; Delaunay, J.; Martinelli, G.; Fiedler, W.; Raffoux, E.; Prebet, T.; Gomez-Roca, C.; Papayannidis, C.; Kebenko, M.; Paschka, P.; et al. Phase I clinical study of RG7356, an anti-CD44 humanized antibody, in patients with acute myeloid leukemia. Oncotarget 2016, 7, 32532–32542. [Google Scholar] [CrossRef] [Green Version]
  147. Macdonald, J.; Henri, J.; Roy, K.; Hays, E.; Bauer, M.; Veedu, R.N.; Pouliot, N.; Shigdar, S. EpCAM immunotherapy versus specific targeted delivery of drugs. Cancers 2018, 10, 19. [Google Scholar] [CrossRef]
  148. Ragnhammar, P.; Fagerberg, J.; Frödin, J.-E.; Hjelm, A.-L.; Lindemalm, C.; Magnusson, I.; Masucci, G.; Mellstedt, H. Effect of monoclonal antibody 17-1A and gm-CSF in patients with advanced colorectal carcinoma—long-lasting, complete remissions can be induced. Int. J. Cancer 1993, 53, 751–758. [Google Scholar] [CrossRef]
  149. Riethmüller, G.; Schneider-Gädicke, E.; Schlimok, G.; Schmiegel, W.; Raab, R.; Höffken, K.; Gruber, R.; Pichlmaier, H.; Hirche, H.; Pichlmayr, R. Randomised trial of monoclonal antibody for adjuvant therapy of resected Dukes’ C colorectal carcinoma. German Cancer Aid 17-1A Study Group. Lancet 1994, 343, 1177–1183. [Google Scholar] [CrossRef]
  150. Schmoll, H.-J.; Wittig, B.; Arnold, D.; Riera-Knorrenschild, J.; Nitsche, D.; Kroening, H.; Mayer, F.; Andel, J.; Ziebermayr, R.; Scheithauer, W. Maintenance treatment with the immunomodulator MGN1703, a Toll-like receptor 9 (TLR9) agonist, in patients with metastatic colorectal carcinoma and disease control after chemotherapy: A randomised, double-blind, placebo-controlled trial. J. Cancer Res. Clin. Oncol. 2014, 140, 1615–1624. [Google Scholar] [CrossRef]
  151. Fields, A.L.; Keller, A.; Schwartzberg, L.; Bernard, S.; Kardinal, C.; Cohen, A.; Schulz, J.; Eisenberg, P.; Forster, J.; Wissel, P. Adjuvant therapy with the monoclonal antibody edrecolomab plus fluorouracil-based therapy does not improve overall survival of patients with stage III colon cancer. J. Clin. Oncol. 2009, 27, 1941–1947. [Google Scholar] [CrossRef]
  152. Münz, M.; Murr, A.; Kvesic, M.; Rau, D.; Mangold, S.; Pflanz, S.; Lumsden, J.; Volkland, J.; Fagerberg, J.; Riethmüller, G.; et al. Side-by-side analysis of five clinically tested anti-EpCAM monoclonal antibodies. Cancer Cell Int. 2010, 10, 44. [Google Scholar] [CrossRef]
  153. Schmidt, M.; Scheulen, M.E.; Dittrich, C.; Obrist, P.; Marschner, N.; Dirix, L.; Schmidt, M.; Ruttinger, D.; Schuler, M.; Reinhardt, C.; et al. An open-label, randomized phase II study of adecatumumab, a fully human anti-EpCAM antibody, as monotherapy in patients with metastatic breast cancer. Ann. Oncol. Off. J. of Eur. Soc. Med. Oncol. 2010, 21, 275–282. [Google Scholar] [CrossRef]
  154. Gao, A.-G.; Lindberg, F.P.; Finn, M.B.; Blystone, S.D.; Brown, E.J.; Frazier, W.A. Integrin-associated protein is a receptor for the C-terminal domain of thrombospondin. J. Boil. Chem. 1996, 271, 21–24. [Google Scholar] [CrossRef]
  155. Lindberg, F.P.; Bullard, D.C.; Caver, T.E.; Gresham, H.D.; Beaudet, A.L.; Brown, E.J. Decreased resistance to bacterial infection and granulocyte defects in IAP-Deficient mice. Science 1996, 274, 795–798. [Google Scholar] [CrossRef]
  156. Miyashita, M.; Ohnishi, H.; Okazawa, H.; Tomonaga, H.; Hayashi, A.; Fujimoto, T.-T.; Furuya, N.; Matozaki, T. Promotion of neurite and filopodium formation by CD47: Roles of integrins, rac, and Cdc42. Mol. Boil. Cell 2004, 15, 3950–3963. [Google Scholar] [CrossRef]
  157. Liu, Y.; Merlin, D.; Burst, S.L.; Pochet, M.; Madara, J.L.; Parkos, C.A. The role of CD47 in neutrophil transmigration: Increased rate of migration correlates with increased cell surface expression of cd47. J. Boil. Chem. 2001, 276, 40156–40166. [Google Scholar] [CrossRef]
  158. Reinhold, M.I.; Lindberg, F.P.; Kersh, G.J.; Allen, P.M.; Brown, E.J. Costimulation of T cell activation by integrin-associated protein (CD47) is an adhesion-dependent, CD28-independent signaling pathway. J. Exp. Med. 1997, 185, 1–11. [Google Scholar] [CrossRef]
  159. Jaiswal, S.; Jamieson, C.H.; Pang, W.W.; Park, C.Y.; Chao, M.P.; Majeti, R.; Traver, D.; Van Rooijen, N.; Weissman, I.L. CD47 is up-regulated on circulating hematopoietic stem cells and leukemia cells to avoid phagocytosis. Cell 2009, 138, 271–285. [Google Scholar] [CrossRef]
  160. Kim, M.J.; Lee, J.-C.; Lee, J.-J.; Kim, S.; Lee, S.G.; Park, S.-W.; Sung, M.W.; Heo, D.S. Association of CD47 with natural killer cell-mediated cytotoxicity of head-and-neck squamous cell carcinoma lines. Tumor Boil. 2008, 29, 28–34. [Google Scholar] [CrossRef]
  161. Chao, M.P.; Alizadeh, A.A.; Tang, C.; Myklebust, J.H.; Varghese, B.; Gill, S.; Jan, M.; Cha, A.C.; Chan, C.K.; Tan, B.T.; et al. Anti-CD47 antibody synergizes with rituximab to promote phagocytosis and eradicate non-Hodgkin lymphoma. Cell 2010, 142, 699–713. [Google Scholar] [CrossRef]
  162. Chan, K.S.; Espinosa, I.; Chao, M.; Wong, D.; Ailles, L.; Diehn, M.; Gill, H.; Presti, J., Jr.; Chang, H.Y.; van de Rijn, M.; et al. Identification, molecular characterization, clinical prognosis, and therapeutic targeting of human bladder tumor-initiating cells. Proc. Natl. Acad. Sci. USA 2009, 106, 14016–14021. [Google Scholar] [CrossRef] [Green Version]
  163. Manna, P.P.; Frazier, W.A. CD47 mediates killing of breast tumor cells via Gi-dependent inhibition of protein kinase A. Cancer Res. 2004, 64, 1026–1036. [Google Scholar] [CrossRef]
  164. Schulenburg, A.; Blatt, K.; Cerny-Reiterer, S.; Sadovnik, I.; Herrmann, H.; Marian, B.; Grunt, T.W.; Zielinski, C.C.; Valent, P. Cancer stem cells in basic science and in translational oncology: Can we translate into clinical application? J. Hematol. Oncol. 2015, 8, 16. [Google Scholar] [CrossRef]
  165. Fan, G.; Wang, Z.; Hao, M.; Li, J. Bispecific antibodies and their applications. J. Hematol. Oncol. 2015, 8, 130. [Google Scholar] [CrossRef]
  166. Fang, X.; Chen, C.; Xia, F.; Yu, Z.; Zhang, Y.; Zhang, F.; Gu, H.; Wan, J.; Zhang, X.; Weng, W.; et al. CD274 promotes cell cycle entry of leukemia-initiating cells through JNK/Cyclin D2 signaling. J. Hematol. Oncol. 2016, 9, 124. [Google Scholar] [CrossRef]
  167. Zhao, H.; Wang, J.; Kong, X.; Li, E.; Liu, Y.; Du, X.; Kang, Z.; Tang, Y.; Kuang, Y.; Yang, Z.; et al. CD47 promotes tumor invasion and metastasis in non-small cell lung cancer. Sci. Rep. 2016, 6, 29719. [Google Scholar] [CrossRef]
  168. Sikic, B.I.; Lakhani, N.; Patnaik, A.; Shah, S.A.; Chandana, S.R.; Rasco, D.; Colevas, A.D.; O’Rourke, T.; Narayanan, S.; Papadopoulos, K.; et al. First-in-human, first-in-class phase I trial of the anti-CD47 antibody Hu5F9-G4 in patients with advanced cancers. J. Clin. Oncol. 2019, 37, 946–953. [Google Scholar] [CrossRef]
  169. Liu, J.; Wang, L.; Zhao, F.; Tseng, S.; Narayanan, C.; Shura, L.; Willingham, S.; Howard, M.; Prohaska, S.; Volkmer, J.; et al. Pre-clinical development of a humanized anti-CD47 antibody with anti-cancer therapeutic potential. PLoS ONE 2015, 10, e0137345. [Google Scholar] [CrossRef]
  170. Liu, X.; Kwon, H.; Li, Z.; Fu, Y.-X. Is CD47 an innate immune checkpoint for tumor evasion? J. Hematol. Oncol. 2017, 10, 12. [Google Scholar] [CrossRef]
  171. Advani, R.; Flinn, I.; Popplewell, L.; Forero, A.; Bartlett, N.L.; Ghosh, N.; Kline, J.; Roschewski, M.; LaCasce, A.; Collins, G.P.; et al. CD47 blockade by Hu5F9-G4 and rituximab in non-hodgkin’s lymphoma. New Engl. J. Med. 2018, 379, 1711–1721. [Google Scholar] [CrossRef]
  172. Huber, A.; Dammeijer, F.; Aerts, J.G.J.V.; Vroman, H. Current state of dendritic cell-based immunotherapy: Opportunities for in vitro antigen loading of different DC subsets? Front. Immunol. 2018, 9, 2804. [Google Scholar] [CrossRef]
  173. Van Willigen, W.W.; Bloemendal, M.; Gerritsen, W.R.; Schreibelt, G.; De Vries, I.J.M.; Bol, K.F. Dendritic cell cancer therapy: Vaccinating the right patient at the right time. Front. Immunol. 2018, 9, 2265. [Google Scholar] [CrossRef]
  174. Li, J.; Li, W.; Huang, K.; Zhang, Y.; Kupfer, G.; Zhao, Q. Chimeric antigen receptor T cell (CAR-T) immunotherapy for solid tumors: Lessons learned and strategies for moving forward. J. Hematol. Oncol. 2018, 11, 22. [Google Scholar] [CrossRef]
  175. Lin, P.-C.; Hsieh, H.-Y.; Chu, P.-C.; Chen, C.S. Therapeutic opportunities of targeting histone deacetylase isoforms to eradicate cancer stem cells. Int. J. Mol. Sci. 2018, 19, 1939. [Google Scholar] [CrossRef]
  176. Mele, L.; Paino, F.; Papaccio, F.; Regad, T.; Boocock, D.; Stiuso, P.; Lombardi, A.; Liccardo, D.; Aquino, G.; Barbieri, A.; et al. A new inhibitor of glucose-6-phosphate dehydrogenase blocks pentose phosphate pathway and suppresses malignant proliferation and metastasis in vivo. Cell Death Dis. 2018, 9, 572. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of cancer stem cells (CSCs) and their role in chemoresistance. (A) Chemotherapy destroys most of the highly proliferative cancer cells. CSCs are less affected by chemotherapy and can re-initiate cancer and promote cancer progression. (B) CSCs possess several mechanisms that counteract the effect of chemotherapeutic compounds such as the ATP family of transporters, higher expression of the pro-survival factor BCL2 (B-cell lymphoma 2), higher expression of the detoxifying enzymes aldehyde dehydrogenases (ALDHs), a more efficient DNA damage response and a slower rate of cell division or dormancy.
Figure 1. Schematic representation of cancer stem cells (CSCs) and their role in chemoresistance. (A) Chemotherapy destroys most of the highly proliferative cancer cells. CSCs are less affected by chemotherapy and can re-initiate cancer and promote cancer progression. (B) CSCs possess several mechanisms that counteract the effect of chemotherapeutic compounds such as the ATP family of transporters, higher expression of the pro-survival factor BCL2 (B-cell lymphoma 2), higher expression of the detoxifying enzymes aldehyde dehydrogenases (ALDHs), a more efficient DNA damage response and a slower rate of cell division or dormancy.
Cells 08 00926 g001
Figure 2. Schematic representation of the CSC microenvironment (CSC niche). The complex CSC niche contains several cell types including mesenchymal stem cells (MSCs), endothelial cells, cancer associated fibroblasts (CAFs) and immune cells (tumour associated macrophages (TAMs), regulatory T cells (Tregs), myeloid-derived suppressor cells (MDSCs) and T-helper cells (Th17)). These cells secrete different cytokines and growth factors that promote tumourigenesis, tumour progression and immunosuppression.
Figure 2. Schematic representation of the CSC microenvironment (CSC niche). The complex CSC niche contains several cell types including mesenchymal stem cells (MSCs), endothelial cells, cancer associated fibroblasts (CAFs) and immune cells (tumour associated macrophages (TAMs), regulatory T cells (Tregs), myeloid-derived suppressor cells (MDSCs) and T-helper cells (Th17)). These cells secrete different cytokines and growth factors that promote tumourigenesis, tumour progression and immunosuppression.
Cells 08 00926 g002
Table 1. Cell surface markers (solid tumours) that are used for the identification and isolation of CSCs.
Table 1. Cell surface markers (solid tumours) that are used for the identification and isolation of CSCs.
TumourBiomarkersReferences
Breast cancerCD44+/CD24−/low/ALDH+[16]
Prostate cancerCD44+/a2b1+/ALDH+[34]
MelanomaABCB5+[17]
CD20+[18]
CD271+
GlioblastomaCD133+[20]
Colon cancerCD133+/CD44+/ALDH+[21]
EpCAM+/CD44+/CD166+[22]
CD44+/CD24+[23]
CD133+/CD24+[24]
Lung cancerCD133+
CD44+
ALDH+[35,47]
CD117+
Gastric cancerCD133+
CD44+/CD24+[25]
CD90+[26]
CD44+/CD54+
Head and neck cancerCD44+/ALDH+[27]
CD44+/CD66-
Ovarian cancerCD133+[28]
CD44+
ALDH+
CD117+
Pancreatic cancerCD133+/CD44+/CD24+/ESA+[29]
Liver cancerCD133+/CD44+[30]
CD90+[31]
EpCAM+[32]
CD13+[33]
Table 2. Clinically tested strategies that were employed to target CSCs. Two main therapies were used: monoclonal antibodies therapies and dendritic cells-based vaccines.
Table 2. Clinically tested strategies that were employed to target CSCs. Two main therapies were used: monoclonal antibodies therapies and dendritic cells-based vaccines.
TargetTumour TypeTherapyClinical Trial/ReportsResults
CD133Advanced cholangiosarcomaCocktail CD133 CAR-T and CART-EGFRNCT02541370
(Phase I and II)
Toxicity
Glioblastoma multiformeICT-121 DC vaccineNCT02049489
(Phase I)
Not published
CD44Refractory
Acute Myeloid Leukemia
Monoclonal antibody (RG7356)146Limited clinical activity
EpCAMMetastatic colorectal cancerEdrecolomab (Monoclonal antibody)151Limited response
150Not statistically significant
3622W94 (Monoclonal antibody)152Toxicity
ING-1 (Monoclonal antibody) Toxicity
Adecatumumab (Monoclonal antibody)153Favourable response in patients with highest expression
Various types of tumoursEPCAM-targeted CAR-T cellsNCT02729493
(N/A)
Ongoing
NCT02725125
(N/A)
Ongoing
NCT02915445
(Phase I)
Ongoing
NCT03013712
(Phase I and II)
Ongoing
CD47Various solid tumours and
hematological malignancies
Hu5F9-G4 (Monoclonal antibody)NCT02216409
(Phase I)
Not statistically significant
NCT02678338
(Phase I)
Not published
Combination Hu5F9-G4 and CetuximabNCT02953782
(Phase I and II)
Ongoing
Combination Hu5F9-G4 and RetuximabNCT02953509
(Phase I and II)
Ongoing
CC-90002 (Monoclonal antibody)NCT02641002
(Phase I)
Ongoing
Combination CC-90002 and RetuximabNCT02367196
(Phase I)
Ongoing
Autogenic glioma
stem-like
cells
(A2B5+)
Glioblastoma multiformeDendritic cell-based vaccineNCT01567202
(Phase II)
Ongoing

Share and Cite

MDPI and ACS Style

Barbato, L.; Bocchetti, M.; Di Biase, A.; Regad, T. Cancer Stem Cells and Targeting Strategies. Cells 2019, 8, 926. https://doi.org/10.3390/cells8080926

AMA Style

Barbato L, Bocchetti M, Di Biase A, Regad T. Cancer Stem Cells and Targeting Strategies. Cells. 2019; 8(8):926. https://doi.org/10.3390/cells8080926

Chicago/Turabian Style

Barbato, Luisa, Marco Bocchetti, Anna Di Biase, and Tarik Regad. 2019. "Cancer Stem Cells and Targeting Strategies" Cells 8, no. 8: 926. https://doi.org/10.3390/cells8080926

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop